Sunteți pe pagina 1din 13

Proceedings of IMECE 2005

ASME 2005 International Mechanical Engineering &


Congress and Exposition
November 5-11, 2005, Orlando Florida, USA
IMECE 2005-79941
FLUID FLOW AND HEAT TRANSFER IN POWER-LAW FLUIDS ACROSS CIRCULAR
CYLINDERS - ANALYTICAL STUDY
Waqar A. Khan, Richard J. Culham, Milan M. Yovanovich
Microelectronics Heat Transfer Laboratory
Department of Mechanical Engineering
University of Waterloo, Waterloo, Ontario, Canada N2L 3G1
Email: wkhan@mhtlab.uwaterloo.ca
ABSTRACT
An integral approach of the boundary layer analysis is em-
ployed for the modeling of uid ow around and heat transfer
from innite circular cylinders in power-law uids. The Von
Karman-Pohlhausenmethod is used to solve the momentuminte-
gral equation whereas the energy integral equation is solved for
both isothermal and isoux boundary conditions. A fourth-order
velocity prole in the hydrodynamic boundary layer and a third-
order temperature prole in the thermal boundary layer are used
to solve both integral equations. Closed form expressions are
obtained for the drag and heat transfer coefcients that can be
used for a wide range of the power-law index, and generalized
Reynolds and Prandtl numbers. It is found that pseudoplastic
uids offer less skin friction and higher heat transfer coefcients
than dilatant uids. As a result, the drag coefcients decrease
and the heat transfer increases with the decrease in power-law
index. Comparison of the analytical models with available ex-
perimental/numerical data proves the applicability of the inte-
gral approach for power-law uids.
NOMENCLATURE
C
D
total drag coefcient
C
Df
friction drag coefcient
C
Dp
pressure drag coefcient
C
f
skin friction coefcient 2
w
/U
2
app
C
p
pressure coefcient 2P/U
2
app
c
p
specic heat of the uid [J/kg K]
D cylinder diameter [m]
k thermal conductivity [W/m K]
h average heat transfer coefcient [W/m
2
K]
m consistency index for non-Newtonian
viscosity [Pa s
n
]
n power-law index
Nu
D
average Nusselt number based on the diameter
of the cylinder hD/k
f
Pr
p
Prandtl number for power-law uids
(U
app
D/)Re
2/(n+1)
Dp
P pressure [N/m
2
]
q heat ux [W/m
2
]
Re
Dp
generalized Reynolds number based on the
diameter of the cylinder D
n
U
2n
app
/m
s distance along the curved surface of the
circular cylinder measured from the forward
stagnation point [m]
T temperature [

C]
U
app
approach velocity [m/s]
U(s) potential ow velocity just outside the boundary
layer 2U
app
sin [m/s]
u s - component of velocity in the boundary layer
[m/s]
v - component of velocity in the boundary layer
[m/s]
1 Copyright c 2005 by ASME
Greek Symbols
thermal diffusivity [m
2
/s]
hydrodynamic boundary-layer thickness [m]

1
displacement thickness [m]

2
momentum thickness [m]

T
thermal boundary layer thickness [m]
distance normal to and measured from the surface
of the circular cylinder [m]
pressure gradient parameter
absolute viscosity of the uid [Ns/m
2
]
kinematic viscosity of the uid [m
2
/s]
density of the uid [kg/m
3
]
shear stress [N/m
2
]
angle measured from front stagnation point [radian]
ratio of thermal and hydrodynamic boundary
layers
T
/
subscripts
a ambient
f uid or friction
H hydrodynamic
p pressure
s separation
T thermal or temperature
w wall
INTRODUCTION
Many practical situations need a knowledge of uid ow
around and heat transfer from horizontal cylinders subjected to
crossow of non-Newtonian uids. These uids are classied by
different authors in different ways. One important classication
is the purely viscous uids (Cho [1]). These uids are further
subdivided in to shear thinning (or pseudoplastic, e. g. paints,
glues, blood, and suspensions) and shear thickening (or dilatant,
e. g. wet sand, sugar and borax solutions) uids. In this study,
uid ow and heat transfer characteristics of these uids across a
circular cylinder are investigated using a power-law model. This
model is based on the fact that both uids exhibit a region of lin-
ear relationship between stress and strain rate when viewed on
a log-log plot (Chhabra and Richardson [2]). This is why, pure
viscous uid are also called power-law uids.
Literature Review
Fluid ow around and heat transfer from circular cylin-
ders in crossow to Newtonian uids has been extensively stud-
ied theoretically, experimentally and numerically by many re-
searchers (Khan et al. [3]) but for non-Newtonian and especially
for power-law uids these types of studies are very limited. From
a theoretical point of view, Acrivos et al. [4] were the rst who
investigated the forced convective heat transfer from an isother-
mal at plate to power-law uids using analytical-numerical ap-
proach. Later, Schowalter [5], Shah et al. [6], Acrivos et al. [7],
and Lee and Ames [8] extended that work to solve 2-D boundary
layer equations by using similarity transformations. Bizzell and
Slattery [9] used the Von Karman-Pohlhausen integral method
to analyze boundary layer on a sphere and calculated points of
separation for different values of power-law index. However, no
attempt was made to investigate heat transfer to power-law u-
ids. Wolf and Szewczyk [10] used Blasius series approach to
investigate heat transfer from arbitrary cylinders to power-law
uids while Serth and Kiser [11] used Goertler series method to
solve 2-D boundary layer equations for power-law uids. Lin
and Chern [12], and Kim et al. [13] presented laminar momen-
tum boundary-layer analyses for non-Newtonian uids by using
the Merk-Chao expansion method. They obtained numerical and
closed form solutions to the universal functions and then applied
themto analyze the wedge owand the owover a circular cylin-
der and a sphere.
Mizushina and Usui [14] developed a theoretical framework for
the laminar boundary heat transfer from a horizontal cylinder
submerged in power-law uids and expressed Nusselt number
at the front stagnation point in the form
Nu
D
= 1.04n
0.4
Re
1/(n+1)
Dp
Pr
1/3
p
(1)
Nakayama et al. [15], Shenoy and Nakayama [16], Nakayama
[17], and Anderson [18] extended the Von Karman-Pohlhausen
integral method to obtain solutions to both momentum and en-
ergy equations for power-law uid ows past wedges, cones and
axisymmetric bodies.
From experimental point of view, Shah et al. [6], Luikov
et al. [19-21], James and Acosta [22], Takahashi et al. [23],
Mizushina et al. [24], Mizushina and Usui [14], Kumar et al.
[25], Ghosh et al. [26], Rao [27] measured local heat/mass
transfer coefcients for the non-Newtonian uids from a circu-
lar cylinder in crossow and determined average heat transfer.
Mizushina et al. [24] developed the following empirical correla-
tion for the Nusselt number averaged over the circumference of
the cylinder:
Nu
D
= 0.72n
0.4
Re
1/(n+1)
Dp
Pr
1/3
p
(2)
whereas Ghosh et al. [26] developed the followingempirical cor-
relation for heat and mass transfer from a cylinder in crossow
to power-law uids:
Nu
D
= 0.785Re
1/2
Dp
Pr
1/3
p
for 10 < Re
Dp
< 25000 (3)
Coelho and Pinho [28-30] performed a series of experiments to
study the ow of non-Newtonian uids around a cylinder. They
2 Copyright c 2005 by ASME
measured shedding frequency, the formation length (l
f
) and the
pressure distribution around a cylinder and determined the shed-
ding regimes and the drag coefcients.
DAlessio and Pascal [31] investigated numerically the
steady power-law ow around a circular cylinder at three dif-
ferent Reynolds numbers Re
Dp
= 5, 20, and 40 using a rst-
order accurate difference method for a xed blockage ratio. They
found that the critical Reynolds number, wake length, separa-
tion angle and drag coefcient depend on the power-law index.
Chhabra et al. [32] extended that work by using a more accu-
rate second-order nite difference method, more rened compu-
tational meshes, and greater blockage ratio and power-law index
ranges in order to investigate the effect of blockage on drag co-
efcient, wake length, separation angle, and ow patterns over
wide ranges of conditions. Agarwal et al. [33] investigated nu-
merically the momentum and thermal boundary layers for power-
law uids over a thin needle under wide ranges of kinematic and
physical conditions. They introduced a similarity variable and
transformed the momentum and energy equations into ordinary
differential equations. They reported extensive results on axial
velocity proles, shear stress and skin friction distributionon the
surface of the needle, temperature and local Nusselt number vari-
ation, and momentum and thermal boundary layer thicknesses in
the followingranges of conditions: 0.2 n 1.6, 1 Pr 1000,
Re < 10
6
and for three needle sizes.
In this study, an approximate method, based on the Karman-
Pohlhausen integral momentum and energy equations, is used to
study the uid ow and heat transfer in power-law uids across
a single circular cylinder.
ANALYSIS
Consider a uniform ow of a non-Newtonian (power-law)
uid past a xed circular cylinder of diameter D, with vanishing
circulation around it, as shown in Fig. 1. The approaching veloc-
ity of the uid is U
app
and the ambient temperature is assumed to
be T
a
. The surface temperature of the wall is T
w
(>T
a
) in the case
of the isothermal cylinder and the heat ux is q for the isoux
boundary condition. The ow is assumed to be laminar, steady,
and two dimensional. The uid is assumed to be incompressible
with constant thermophysical and rheological properties. The
potential ow velocity just outside the boundary layer is denoted
by U(s). Using an order-of-magnitude analysis (Khan [34]), the
reduced equations of continuity, momentum and energy in the
curvilinear system of coordinates (Fig. 1) for a power-law uid
can be written as:
Continuity:
u
s
+
v

= 0 (4)

A
D

s
s U
a app
T U
s

x
v

S
Figure 1. Flow over a circular cylinder.
s-Momentum:
u
u
s
+v
u

=
1

dP
ds
+
1

(5)
-Momentum:
dP
d
= 0 (6)
Energy:
u
T
s
+v
T

2
T

2
(7)
with

dP
ds
=U(s)
dU(s)
ds
(8)
and

w
= m
_
u

_
n

=0
(9)
where m is a consistency index for non-Newtonian viscosity and
n is called power-law index, that is < 1 for pseudoplastic, =1 for
Newtonian, and > 1 for dilatant uids.
3 Copyright c 2005 by ASME
Hydrodynamic Boundary Conditions
The assumptions of no slip boundary condition at the cylin-
der wall (u =0, at = 0), no mass ow through the cylinder wall
(v = 0 at = 0), and the potential ow just outside the boundary
layer [u = U(s) at = (s)] give the following complete set of
hydrodynamic conditions:
(i) at the cylinder surface, i.e., at = 0
u = 0 and

2
u

2
=
U(s)
dU(s)
ds
n
_
u

_
n1
(10)
where = m/.
(ii) at the edge of the boundary layer, i.e., at = (s)
u =U(s),
u

= 0 and

2
u

2
= 0 (11)
These conditions will help in determining the velocity distribu-
tion inside the boundary layer.
Thermal Boundary Conditions
The assumptions of uniform wall temperature (UWT) and
uniformwall ux (UWF) boundary conditions give the following
complete set thermal conditions:
(i) at the cylinder surface, i.e., at = 0

2
T

2
= 0 and
_

_
T = T
w
for UWT
T

=
q
k
f
for UWF
(12)
(ii) at the edge of thermal boundary layer, i.e., at =
T
T = T
a
and
T

= 0 (13)
Using these thermal conditions, the temperature distributions in-
side the thermal boundary layer can be determined.
Velocity Distribution
Assuming a thin hydrodynamic boundary layer around the
cylinder, the velocity distribution inside the boundary layer, sat-
isfyingEqs. (10) and (11), can be approximated by a fourth order
polynomial as suggested by Pohlhausen [35]:
u
U(s)
= (2
H
2
3
H
+
4
H
) +

6
(
H
3
2
H
+3
3
H

4
H
) (14)
where 0
H
= /(s) 1 and is the pressure gradient pa-
rameter, given by
=
dU(s)
ds

n+1
U(s)
1n
_
2 +

6
_
1n
n
(15)
With the help of velocity proles, Schlichting [36] showed that
the parameter is restricted to the range 12 12.
Temperature Distribution
Assuming a thin thermal boundary layer around the cylin-
der, the temperature distribution in the thermal boundary layer,
satisfying Eqs. (12) and (13), can be approximated by a third
order polynomial
T T
a
T
w
T
a
= 1
3
2

T
+
1
2

3
T
(16)
for the isothermal boundary condition and
T T
a
=
2q
T
3k
f
_
1
3
2

T
+
1
2

3
T
_
(17)
for the isoux boundary condition.
Boundary Layer Parameters
In dimensionless form, the momentum integral equation can
be written as
U
2
d
2
ds
+(2
2
+
1
)U
dU
ds
=
1


w
(18)
where
1
and
2
are the displacement and momentum boundary
layer thicknesses and are given by

1
=

1
0
_
1
u
U(s)
_
d (19)
and

2
=

1
0
u
U(s)
_
1
u
U(s)
_
d (20)
4 Copyright c 2005 by ASME
Using velocity distributionfrom Eq. (14), Eqs. (19) and (20) can
be written as:

1
=

10
_
3

12
_
(21)
and

2
=

63
_
37
5


15


2
144
_
(22)
Simplifying and arranging the terms in Eq. (18), we get
U
n
2
U
n1
d
2
ds
+
_
2 +

1

2
_

n+1
2
U
1n

dU
ds
=
_

_
n
_
2 +

6
_
n
(23)
Assuming
Z =

2
n+1
U
1n

and K = Z
dU
ds
Equation (23) can be reduced to a non-linear differential equation
of the rst order for Z, which can be written as:
dZ
ds
=
F
U
(24)
where
F =
K
_

_
(n +1)
_
2 +

6
_
n
(1 +3n)
_

_
(n +1)
_

_
_

_
(25)
At the stagnation point s = 0, U = 0. Since dZ/ds cannot be
innite, F must be zero at the stagnation point. Hence
(n +1)
_
2 +

6
_
n
(1+3n)
_

_
(n+1)
_

_
=0 (26)
which gives the values of the pressure gradient parameter for
different values of n at the stagnation point. Due to limitations
of the method used in this study, no root of Eq. (26) could
be found in the range 12 12 for n < 0.895. Bizzell
and Slattery [9] could calculate the roots for 0.7358 n 1.0
only whereas Mizushina and Usui [14] calculated the roots in
the range 0.895 n 1.19. In the present study, the values of
are calculated for n 0.895. These values are plotted in Fig. 2
as a function of n.
Following Walz [37], the function F can be approximated
by a straight line
F = a bK (27)
where the constants a and b are determined for each power index
n and are correlated in to single correlations
a = 0.45n
1.25
and b = 6.2n
0.8
(28)
These correlations are valid between the stagnation point (F =0)
and the point of maximum velocity (K = 0). So Eq. (24) can be
written as
U(s)
dZ
ds
= a bK (29)
Using potential owvelocity outside the boundary layer for a cir-
cular cylinder and rearranging the terms, Eq. (29) can be solved
for the local dimensionless momentum thickness:

2
D
=
_
a(sin)
1+bn
2
3n
Re
Dp


0
(sin)
b1
d
_
1/(n+1)
(30)
where is a dummy variable and Re
Dp
is the generalized
Reynolds number for a power-law uid and is dened as
Re
Dp
=
D
n
U
2n
app
m
(31)
From Eq. (15), the dimensionless hydrodynamic boundary layer
thickness can be written as

D
=
_

_
1
Re
Dp

nsin
n

2
2n
sin2
_
2 +

6
_
1n
_

_
1/(n+1)
(32)
Solving Eq. (22) with Eq. (32) and comparing the results with
the numerical values obtained from Eq. (30), one can get the val-
ues of pressure gradient parameter for each position along the
5 Copyright c 2005 by ASME
(Radians)

0.5 1 1.5 2
-12
-6
0
6
12
0.8943
1
1.2
1.4
n
Figure 2. Behavior of Pressure Gradient Parameter for Power-Law Flu-
ids.
n

s
(
d
e
g
r
e
e
s
)
0.4 0.8 1.2 1.6 2
85
90
95
100
105
110
115
Pseudoplastic Fluids Dilatant Fluids
N
e
w
t
o
n
i
a
n
F
l
u
i
d
s
Figure 3. Angles of Separation for Power-Law Fluids.
circumference of the cylinder. These values were tted by the
least squares method for each power-law index n and are shown
in Fig. 2 for pseudoplastic and dilatant uids. It shows that the
value of decreases with increasing n. Using analytical deni-
tion of the point of separation, the angle of separation was deter-
mined which depends on the nature of the uid (pseudoplastic or
dilatant) as well as on the velocity distribution inside the bound-
ary layer. The calculated angles of separation are plotted in Fig.
3 as a function of n. It shows that the separation angle decreases
with the increase in power-law index n,which is in accordance
with Serth and Kiser [11].
Fluid Flow
The rst parameter of interest is uid friction which mani-
fests itself in the form of the drag force F
D
, where F
D
is the sum
of the skin friction drag D
f
and pressure drag D
p
. Skin friction
drag is due to viscous shear forces produced at the cylinder sur-
face, predominantly in those regions where the boundary layer is
attached. In dimensionless form, it can be written as
C
f
=

w
1
2
U
2
app
(33)
Using Eqs. (9) and (14) and simplifying, we get
C
f
=
2
Re
1/(n+1)
Dp
_
(+12) sin
3
_
n
_
sin2
2
n2
(2 +

6
)
1n
nsin
n

_
n/(n+1)
(34)
The friction drag coefcient can be dened as
C
Df
=


0
C
f
sind
=


s
0
C
f
sind+

s
C
f
sind (35)
Since the shear stress on the cylinder surface after boundary layer
separation is very small, the second integral can be neglected and
the friction drag coefcient can be written as
C
Df
=


s
0
C
f
sind (36)
The calculations were performed for different values of n and the
results are correlated in terms of n and the generalized Reynolds
number Re
Dp
in to a single correlation
C
Df
=
5.786n
0.32
Re
1/(n+1)
Dp
(37)
which gives friction drag coefcient for the ow of a power-law
uid over a circular cylinder in an innite medium. It is interest-
ing to note here that for n = 1, Eq. (37) gives the friction drag
coefcient for the ow of Newtonian uid over a circular cylin-
der (Khan, et al. [3]).
Pressure drag is due to the unbalanced pressures which exist
between the relatively high pressures on the upstream surfaces
and the lower pressures on the downstream surfaces. In dimen-
sionless form, it can be written as
C
Dp
=


0
C
p
cosd (38)
6 Copyright c 2005 by ASME
where C
p
is the pressure coefcient and can be dened as
C
p
=
P
1
2
U
2
app
(39)
The pressure difference P can be obtained by integrating -
momentum equation with respect to . Following Shibu et al.
[38], the -momentum equation, for power-law uids, can be
written as
u
r
u

r
+
u

r
u

+
u
r
u

r
=
1
2r
C
p

+
2
n
Re
Dp
(40)

_
1
r
2

r
(r
2

r
) +
1
r

_
where

r
= 2
r

= 2

with
= (2)
(n1)/2

r
=
1
2
_
r

r
_
u

r
_
+
1
r
u
r

=
1
r
u

+
u
r
r
=
2
rr
+
2

+2
2
r
and
u
r
= cos(1
1
r
2
) u

=sin(1 +
1
r
2
)
In Eq. (40), Shibu et al. [38] scaled the velocity terms us-
ing U
app
, pressure with
1
2
U
2
app
, radial coordinate by the radius
of the cylinder R, the stress components by m(U
app
/R)
n
, and
the second invariant of the rate of deformation tensor, using
(U
app
/R)
2
. Using derivatives of the velocity components and
an order-of-magnitude analysis (Khan [34]), Eq. (40) can be re-
duced to
C
p

= 4sin2
2
3n
Re
Dp
sin (41)
Integrating it with respect to , we get
C
p
= 2(1 cos2) +
2
3n
Re
Dp
(1 cos) (42)
So, the pressure drag coefcient for the cylinder up to the sepa-
ration point will be
C
Dp
=


s
0
C
p
cosd (43)
This pressure drag coefcient was calculated for different values
of n and correlated in to a single correlation in terms of n and
Re
Dp
:
C
Dp
=
1.26n
3.25
Re
Dp
+1.28[1 exp(2.4n)] (44)
The total drag coefcient C
D
can be written as the sum of both
drag coefcients
C
D
=
5.786n
0.32
Re
1/(n+1)
Dp
+
1.26n
3.25
Re
Dp
+1.28[1 exp(2.4n)] (45)
which agrees with the drag coefcient of a Newtonian uid (n =
1) over a circular cylinder (Khan, et al. [3]).
Heat Transfer
The second parameter of interest is the dimensionless average
heat transfer coefcient, Nu
D
for large Prandtl numbers. This
parameter is determined by integrating Eq. (7) from the cylin-
der surface to the thermal boundary layer edge. Assuming the
presence of a thin thermal boundary layer
T
along the cylinder
surface, the energy integral equation for the isothermal boundary
condition can be written as
d
ds


T
0
(T T
a
)ud =
T

=0
(46)
Using velocity and temperature proles Eqs. (14) and (16), and
assuming =
T
/ < 1, Eq. (46) can be simplied to

T
d
ds
[U(s)
T
(+12)] = 90 (47)
This equation can be rewritten separately for the two regions
(Fig. 1), i.e.

T
d
ds
[U(s)
T
(
1
+12)] = 90 (48)
for the region I, and

T
d
ds
[U(s)
T
(
2
+12)] = 90 (49)
for the region II. Integrating Eqs. (48) and (49), in the respective
regions, with respect to s, one can obtain local thermal boundary
7 Copyright c 2005 by ASME
layer thicknesses
_

T
()
D
_
Re
1/(n+1)
Dp
Pr
1/3
p
=
_

_
3
_
45F
1
(n, ) for region I
3
_
45F
2
(n, ) for region II
(50)
where the Prandtl number, Pr
p
, for a power-law uid is dened
as
Pr
p
=
U
app
D

Re
2/(n+1)
Dp
(51)
The functions F
1
(n, ) and F
2
(n, ) in Eq. (50) are given by
F
1
(n, ) =
f
1
()
sin
2
(
1
+12)
2
_

_
2
n2
n
1
sin
n

sin2
_

1
+12
6
_
1n
_

_
1
n +1
F
2
(n, ) =
f
3
()
sin
2

_
2
n2
n
2
sin
n

sin2
_

2
+12
6
_
1n
_

_
1/(n+1)
_

_
(52)
with
f
1
() =


0
sin(
1
+12)d
f
2
() =


s

1
sin(
2
+12)d
f
3
() =
f
1
()

1
+12
+
f
2
()

2
+12
_

_
(53)
The local heat transfer coefcients, for the isothermal boundary
condition, in both the regions can be written as
h
1
() =
3k
f
2
T
1
and h
2
() =
3k
f
2
T
2
(54)
Thus the dimensionless local heat transfer coefcients, for both
the regions, can be written as
Nu
D
()|
isothermal
Re
1/(n+1)
Dp
Pr
1/3
p
=
3
2
_

_
3
_
1
45F
1
(n, )
for region I
3
_
1
45F
2
(n, )
for region II
(55)
The average heat transfer coefcient is dened as
h =
1


0
h() d
=
1


s
0
h() d+

s
h() d
_
(56)
The integral analysis is unable to predict local heat transfer val-
ues from separation point to the rear stagnation point (second
integral on R.H.S.). The experiments (

Zukauskas and

Ziug zda
[39], Fand and Keswani [40], and Nakamura and Igarashi [41]
among others for Newtonian uids and Rao [27] for Non-
Newtonian uids) show that the heat transfer from the rear por-
tion of the cylinder increases with Reynolds numbers. For New-
tonian uids, Van der Hegge Zijnen [42] demonstrated that the
average heat transferred from the rear portion of the cylinder can
be determined from Nu
D
= 0.001Re
D
that shows the weak de-
pendence of average heat transfer from the rear portion of the
cylinder on Reynolds numbers. Same weak dependence can be
observed from Rao [27] experiments for non-Newtonian uids.
In order to include the share of heat transfer from the rear portion
of the cylinder, the local heat transfer coefcients are integrated
upto the separation point and averaged over the whole surface,
that is
h =
1


s
0
h() d
=
1


1
0
h
1
() d+

1
h
2
() d
_
(57)
Using Eqs. (52) - (54), Eq. (57) was solved for different values
of n and the results were correlated in to a single correlation as a
function of n, Re
Dp
, and Pr
p
:
Nu
D
|
isothermal
= 0.593n
0.17
Re
1/(n+1)
Dp
Pr
1/3
p
(58)
For the isoux boundary condition, the energy integral equation
can be written as
d
ds


T
0
(T T
a
)ud =
q
c
p
(59)
Assuming constant heat ux and thermophysical properties and
using Eqs. (14) and (17), Eq. (59) can be simplied to
d
ds
_
U(s)
2
T
(+12)

= 90
k
f
c
p
(60)
8 Copyright c 2005 by ASME
Rewriting Eq. (60) for the two regions in the same way as Eq.
(47), one can obtain local thermal boundary layer thicknesses

T
1
and
T
2
under isoux boundary condition. The local surface
temperatures for the two regions can then be obtained from tem-
perature distribution
T
1
() =
2q
T
1
3k
f
and T
2
() =
2q
T
2
3k
f
(61)
The local heat transfer coefcient can now be obtained from its
denition as
h
1
() =
q
T
1
()
and h
2
() =
q
T
2
()
(62)
which give the local Nusselt numbers for the cross ow over a
cylinder with constant ux
Nu
D
()|
iso f lux
Re
1/(n+1)
Dp
Pr
1/3
=
3
2
_

_
3
_
2
45G
1
(n, )
for region I
3
_
2
45G
2
(n, )
for region II
(63)
where
G
1
(n, ) =

sin(
1
+12)
_

_
2
n2
n
1
sin
n

sin2
_

1
+12
6
_
1n
_

_
1
n +1
G
2
(n, ) =
g()
sin
_

_
2
n2
n
2
sin
n

sin2
_

2
+12
6
_
1n
_

_
1
n +1
_

_
(64)
with
g() =
_

1
+12
+
/2

2
+12
_
(65)
Following the same procedure for the average heat transfer co-
efcient as mentioned above, one can obtain the average Nusselt
number for an isoux cylinder as
Nu
D
|
iso f lux
= 0.627n
0.19
Re
1/(n+1)
Dp
Pr
1/3
p
(66)
Combining the results for both thermal boundary conditions, we
have
Nu
D
Re
1/(n+1)
Dp
Pr
1/3
p
=
_
0.593n
0.17
for UWT
0.627n
0.19
for UWF
(67)
These correlations agree with the heat transfer coefcients for
a Newtonian uid (n = 1) over a circular cylinder (Khan, et al.
[3]).
RESULTS AND DISCUSSION
Flow Characteristics
The dimensionless local shear stress, C
f
, is plotted in Fig.
4 for different power-law uids. It shows that C
f
is zero at the
stagnation point for each uid and reaches a maximum at
58

.
(rad)
C
f
0.5 1 1.5 2 2.5
0
0.15
0.3
0.45
0.8943
1.0
1.2
1.4
Numerical
n
Re
Dp
= 1000
P
s
e
u
d
o
p
l
a
s
t
i
c
F
l
u
i
d
s
D
i
l
a
t
a
n
t
F
l
u
i
d
s
Newtonian Fluids
Schonauer (1964)
(n = 1)
Figure 4. Effect of Power Index n on Skin Friction for a Circular Cylinder.
The increase in shear stress is caused by the deformation of
the velocity proles in the boundary layer, a higher velocity gra-
dient at the wall and a thicker boundary layer. In the region of de-
creasing C
f
preceeding the separation point, the pressure gradi-
ent decreases further and nally C
f
falls close to zero around the
separation point, where boundary-layer separation occurs. Be-
yond this point, C
f
remains close to zero up to the rear stagnation
point. It also shows that the skin friction increases with the in-
crease in power-lawindex n. Thus pseudoplastic (shear-thinning)
uids offer less skin friction than dilatant (shear-thickening) u-
ids which is in accordance with the numerical results of Agarwal
et. al [33]. The results of Newtonian uids (n = 1) are compared
9 Copyright c 2005 by ASME
with the numerical data of Sch onauer [43], which shows good
agreement for the entire range.
n
C
D
1 1.2 1.4 1.6
1.5
2
2.5
3
R
e D
p
=
5
0
R
eD
p
=
1
0
0
Figure 5. Effect of n on Drag Coefcients for a Circular Cylinder.
The variation of the total drag coefcient C
D
with the power-
law index n for different Reynolds numbers is illustrated in Fig.
5. It shows that for a given Reynolds number, the drag coefcient
C
D
increases linearly with n and for a given uid it decreases
with the increase in Reynolds number. The drag coefcients for
a pseudoplastic uids are found to be lower than dilatant uids.
+
+
+
+
+
+
+
+
+
+
+
+
Re
Dp
C
D
10
0
10
1
10
2
10
3 10
-1
10
0
10
1
10
2
Experimental, n=1 (Wieselsberger, 1921)
n=0.6
n=1
n=1.6
+
}
Present Model
Figure 6. Effect of Re
Dp
on Drag Coefcients for a Circular Cylinder.
The effect of Reynolds number on the drag coefcients for
different uids is shown in Fig. 6. Since there are no other exper-
imental/numerical results to compare with pseudoplastic/dilatant
uids, comparisons are made with the experimental results of
Wieselsberger for the air (n = 1) only. The comparison shows
good agreement with the non-Newtonian case.
Heat Transfer Characteristics
The heat transfer parameter (HTP) Nu
D
/Re
1/(n+1)
Dp
Pr
1/3
p
is
presented in Fig. 7 for both the isothermal and isoux boundary
conditions . It shows that HTP decreases with the increase in
the power law index n. Thus pseudoplastic uids transfer more
heat than dilatant uids for the same thermal boundary condi-
tion. The isoux boundary condition gives a higher heat transfer
coefcient for both types of uids.
n
N
u
D
/
R
e
D
p 1
/
n
+
1
P
r
p 1
/
3
0 0.5 1 1.5 2
0.4
0.5
0.6
0.7
0.8
0.9
UWF
UWT
Dilatant Fluids
N
e
w
t
o
n
i
a
n
F
l
u
i
d
s
Pseudoplastic
Fluids
Figure 7. Comparison of Heat Transfer Parameters for Isothermal and
Isoux Thermal Boundary Conditions.
The average heat transfer coefcients Nu
D
/Pr
1/3
p
versus
Re
Dp
are presented in Fig. 8 for different uids. Experimental
results of Takahashi et al. (1977) and Mizushina et al. (1978) for
different CMC (Carboxy Methyl Cellulose) solutions are com-
pared with the present model for isothermal boundary condition.
The comparison is found to be in good agreement for all uids.
Figure 9 shows the comparison of the average heat transfer
coefcients Nu
D
/Pr
1/3
p
versus Re
Dp
for Newtonian uids. Here,
the experimental results of Hilpert (1933) and McAdams (1939)
for air are compared with the present model for n = 1. The re-
sults are found to be in good agreement for the entire range of
Reynolds numbers.
10 Copyright c 2005 by ASME
Re
Dp
N
u
D
/
P
r
p 1
/
3
10
1
10
2
10
3
10
4
10
5 10
0
10
1
10
2
10
3
Present Model
Experimental (Mizushina et. al,1978)
Experimental (Mizushina et. al,1978)
Present Model
Takahashi et al., 1977 (n=0.914)
}
}
n = 0.78
n = 0.93
Figure 8. Comparison of Heat Transfer Coefcients from an Isothermal
Circular Cylinder to CMC Solutions.
Re
Dp
N
u
D
/
P
r
p 1
/
3
10
2
10
3
10
4 10
0
10
1
10
2
10
3
n
=
0
.5
0
.8
1
.2
1
.4
1
.6
1
.0
Hilpert, 1933
McAdams, 1939
Experimental
n = 1
}
Figure 9. Comparison of Heat Transfer Coefcients for a Circular Cylin-
der.
CONCLUSIONS
An integral approach is employed to investigate the uid
ow and heat transfer from an isolated circular cylinder sub-
merged in power-lawuids. Closed formsolutions are developed
for both the drag and heat transfer coefcients in terms of gen-
eralized Reynolds and Prandtl numbers. The correlations of heat
transfer are developed for both isothermal and isoux boundary
conditions. It is found that pseudoplastic uids offer less skin
friction and higher heat transfer coefcients than dilatant uids.
Furthermore, the drag coefcients decrease and the heat transfer
increases with the decrease in power-law index. It is shown that
the present results are in good agreement with the experimen-
tal results for the full laminar range of Reynolds numbers in the
absence of free stream turbulence and blockage effects.
ACKNOWLEDGMENT
The authors gratefully acknowledge the nancial support of
Natural Sciences and Engineering Research Council of Canada
and the Center for Microelectronics Assembly and Packaging.
REFERENCES
[1] Cho, Y. I. and Hartnett, J. P., 1985, Non-Newtonian Flu-
ids, Handbook of Heat Transfer Applications, Second edi-
tion, Chapter 2, McGraw-Hill Book Company, New York.
[2] Chhabra, R. P. and Richardson, J. F., 1999, Non-Newtonian
Flowin the Process Industries: Fundamentals and Engineer-
ing Applications, Butterworth-Heinemann, Great Britain.
[3] Khan, W. A., Culham, J. R., and Yovanovich, M. M., Fluid
Flow Around and Heat Transfer From an Innite Circular
Cylinder, ASME Journal of Heat Transfer (in press).
[4] Acrivos, A., Shah, M. J., and Petersen, E. E., 1960, Mo-
mentum and Heat Transfer in Laminar Boundary-Layer
Flows of Non-Newtonian Fluids Past External Surfaces,
A.I.Ch.E. Journal, Vol. 6, No. 2, pp.312-317.
[5] Schowalter, W. R., 1960, Application of Boundary-Layer
Theory to Power-Law Pseudo-Plastic Fluids: Similar Solu-
tions, A.I.Ch.E. Journal, Vol. 6, No. 1, pp.24-28.
[6] Shah, M. J., 1962, Petersen, E. E., and Acrivos, A., Heat
Transfer from a Cylinder to a Power-Law Non-Newtonian
Fluid, A. I. Ch.E. Journal, Vol. 8, No. 4, pp.542-549.
[7] Acrivos, A., Shah, M. J., and Petersen, E. E., 1965, On
Solution of Two-Dimensional Boundary-Layer Flow Equa-
tions for Non-Newtonian Power-Law Fluid, Chemical En-
gineering Science, Vol. 20, No. 2, pp. 101-105.
[8] Lee, S. Y., and Ames, W. F., 1966, Similarity Solutions for
Non-Newtonian Fluids, A.I.Ch.E. Journal, Vol. 12, No. 4,
pp. 700-708.
[9] Bizzle, G. D. and Slattery, J. C., 1962, Non-Newtonian
Boundary-Layer Flow, Chemical Engineering Science, Vol.
17, pp.777-782.
[10] Wolf, C. J. and Szewczyk, A. A., 1966, Laminar Heat
Transfer to Non-Newtonian Fluids from Arbitrary Cylin-
ders, Proceedings of the Third IHTC, Chicago, Illinois, Au-
gust 7-12, Volume 1.
[11] Serth, R. W., and Kiser, K. M., 1967, A Solution
of the Two-Dimensional Boundary-Layer Equations for an
Ostwald-de Waele Fluid, Chemical Engineering Science,
Vol. 22, pp. 945-956.
[12] Lin, F. N. and Chern, S. Y., 1979, Laminar Boundary
ayer Flow of Non-Newtonian Fluid, International Journal
of Heat and Mass Transfer, Vol. 22, pp. 1323-1329.
11 Copyright c 2005 by ASME
[13] Kim, H. W., Jeng, D. R., and DeWitt, K. J., 1983, Mo-
mentum and Heat Transfer in Power-Law Fluid Flow Over
Two-Dimensional or Axisymmetrical Bodies, International
Journal of Heat and Mass Transfer, Vol. 26, No. 2, pp. 245-
259.
[14] Mizushina, T. and Usui, H., 1978, Approximate Solu-
tion of the Boundary Layer Equations for the Flow of a
Non-Newtonian Fluid Around a Cylinder, Heat Transfer:
Japanese Research, Vol. 7, No. 2, pp. 83-92.
[15] Nakayama, A., Shenoy, A. V., and Koyama, H., 1986, An
Analysis for Forced Convection Heat Transfer fromExternal
Surfaces to Non-Newtonian Fluids, Warme-und Stoffuber-
tragung, Vol. 20, pp. 219-227.
[16] Shenoy, A. V., and Nakayam, A., 1986, Forced Con-
vection Heat Transfer fro Axisymmetric Bodies to Non-
Newtonian Fluids, Canadian Journal of Chemical Engi-
neering, Vol. 64, pp. 680-686.
[17] Nakayama, A., 1986, Integral Methods for Forced Con-
vection Heat Transfer in Power-Law Non-Newtonian Flu-
ids, Encyclopedia of Fluid Mechanics: Rheology and Non-
Newtonian Flows, Vol. 7, pp. 305-339, Houston, USA:Gulf.
[18] Anderson, H. I., 1988, The Nakayama-Koyama Approach
to Laminar Forced Convection Heat Transfer to Power-Law
Fluids, International Journal of Heat and Fluid Flow, Vol.
9, No. 3, pp. 343-346.
[19] Luikov, A. V., Schulman, Z. P., and Berkovsky, B. M.,
1966, Heat and Mass Transfer in a Boundary Layer of
Non-Newtonian Fluids, Proceedings of the Third IHTC,
Chicago, Illinois, August 7-12, Volume 1.
[20] Luikov, A. V., Schulman, Z. P., and Puris, B. I.,
1969a, Mass Transfer of Cylinder in Forced Flow of Non-
Newtonian Viscoelastic Fluid, Heat Transfer-Soviet Re-
search, Vol. 1, No. 1, pp. 121-32.
[21] Luikov, A. V., Schulman, Z. P., and Puris, B. I., 1969b,
External Convective Mass Transfer in Non-Newtonian
Fluid, International Journal of Heat and Mass Transfer, Vol.
12, pp. 377-391.
[22] James, D. F., and Acosta, A. J., 1970, The Laminar Flow
of Dilute Polymer Solutions Around Circular Cylinders,
Journal of Fluid Mechanics, Vol. 42, Part 2, pp. 269-288.
[23] Takashi, K., Maeda, M., and Ikai, S., 1977, Experimen-
tal Study of Heat Transfer from a Cylinder Submerged in
a Non-Newtonian Fluid, die dem Nahenungsansatz von K.
Pohlhausen genugen, Lilenthal Bericht 510, p. 335-339.
[24] Mizushina, T., Usui, H., Veno, K., and Kato, T., 1978,
Experiments of Pseudoplastic Fluid Cross Flow Around a
Circular Cylinder, Heat Transfer: Japanese Research, Vol.
7, No. 3, pp. 92-101.
[25] Kumar, S., Mall, B. K., and Upadhyay, S. N., 1980, On
the Mass Transfer in Non-Newtonian Fluids: II Transfer fro
Cylinders to Power-Law Fluids, Letters in Heat and Mass
Transfer, Vol. 7, pp. 55-64.
[26] Ghosh, U. K., Gupta, S. N., Kumar, S., and Upadhay, S.
N., 1986, Mass Transfer in Cross Flow of Non-Newtonian
Fluid Around a Circular Cylinder, International Journal of
Heat and Mass Transfer, Vol. 29, No. 6, pp. 955-960.
[27] Rao, B. K., 2000, Heat Transfer to Non-Newtonian Flows
Over a Cylinder in Cross Flow, International Journal of
Heat and Fluid Flow, Vol. 21, pp. 693-700.
[28] Coelho, P. M., and Pinho, F. T., 2003, Vortex Shedding
in Cylinder Flow of Shear-Thinning Fluids: I. Identica-
tion and Demarcation of Flow Regimes, Journal of Non-
Newtonian Fluid Mechanics, Vol. 110, pp. 143-176.
[29] Coelho, P. M., and Pinho, F. T., 2003, Vortex Shedding in
Cylinder Flow of Shear-Thinning Fluids: II. Flow Charac-
teristics, Journal of Non-Newtonian Fluid Mechanics, Vol.
110, pp. 177-193.
[30] Coelho, P. M., and Pinho, F. T., 2004, Vortex Shedding
in Cylinder Flow of Shear-Thinning Fluids: III. Pressure
Characteristics, Journal of Non-Newtonian Fluid Mechan-
ics, Vol. 121, pp. 55-68.
[31] DAlessio, S. J. D., and Pascal, J. P., 1996, Steady Flowof
a Power-Law Fluid Past a Cylinder, Acta Mechanica, Vol.
117, pp. 87-100.
[32] Chhabra, R. P., Soares, A. A., and Ferreira, J. M., 2004,
Steady Non-Newtonian Flow Past a Circular Cylinder: A
Numerical Study, Acta Mechanica, Vol. 172, pp. 1-16.
[33] Agarwal, M., Chhabra, R. P., and Eswaran, V., 2002,
Laminar Momentum and Thermal Boundary Layers of
Power-Law Fluids Over a Slender Cylinder, Chemical En-
gineering Science, Vol. 57, pp.1331-1341.
[34] Khan, W. A., 2004, Modeling of Fluid Flow and Heat
Transfer for Optimization of Pin-Fin Heat Sinks, Ph. D.
Thesis, Department of Mechanical Engineering, University
of Waterloo, Canada.
[35] Pohlhausen, K., 1921, Zur N aherungsweise Integration
der Differential Gleichung der Laminaren Reibungschicht,
Zeitschrift f ur angewandte Mathematic und Mechanic, Vol.
1, pp. 252-268.
[36] Schlichting, H., 1979, Boundary Layer Theory, 7th Edi-
tion., McGraw-Hill, New York.
[37] Walz, A., 1941, Ein neuer Ansatz f ur das Greschwindlig-
keitsprol der laminaren Reibungsschicht, Lilienthal-
Bericht 141, p. 8.
[38] Shibu, S., Chhabra, R. P., and Eswaran, V., 2001, Power-
Law Fluid Flow Over a Bundle of Cylinders at Intermediate
Reynolds Numbers, Chemical Engineering Science, Vol.
56, pp. 5545-5554.
[39]

Zukauskas, A. and

Ziug zda, J., 1985, Heat Transfer of a
Cylinder in Crossow, Hemisphere Publishing Corporation,
New York.
[40] Fand, R. M., and Keswani, K. K., 1972, A Continuous
Correlation Equation for Heat Transfer From Cylinders to
Air in Crossow for Reynolds Numbers From 10
2
to 2
12 Copyright c 2005 by ASME
10
5
, International Journal of Heat and Mass Transfer, Vol.
15, pp. 559-562.
[41] Nakamura, H., and Igarashi, T., 2004, Variation of Nusselt
Number with Flow Regimes Behind a Circular Cylinder for
Reynolds Numbers from 70 30000, International Journal of
Heat and Mass Transfer, Vol. 47, pp. 5169-5173.
[42] Van der Hegge Zijnen, B. G., 1956, Modied Corre-
lation Formulae for Heat Transfer by Natural and Forced
Convection from Horizontal Cylinders, Applied Scientic
Research-A, Vol. 6, No. 2-3, pp. 129-140.
[43] Sch onauer, W., 1964, Ein Differenzenverfahren zur
L osung der Grenzschichtgleichung f ur station are, laminare,
inkompressible Str omung, Ing.-Arch., Vol. 33, p. 173.
13 Copyright c 2005 by ASME

S-ar putea să vă placă și